arxiv:math/ v1 [math.dg] 3 Nov 2004

Similar documents
Chapter 2 Noether s First Theorem

arxiv: v1 [math-ph] 2 Apr 2013

Group Actions and Cohomology in the Calculus of Variations

arxiv:gr-qc/ v1 10 Nov 1995

Będlewo. October 19, Glenn Barnich. Physique théorique et mathématique. Université Libre de Bruxelles & International Solvay Institutes

Deformations of coisotropic submanifolds in symplectic geometry

arxiv:math-ph/ v3 13 May 2005

Higher-Spin Fermionic Gauge Fields & Their Electromagnetic Coupling

Invariant Calculus with Vessiot

BACKGROUND IN SYMPLECTIC GEOMETRY

Higher-Spin Fermionic Gauge Fields & Their Electromagnetic Coupling

Iterated Differential Forms III: Integral Calculus

arxiv: v1 [math-ph] 13 Oct 2015

CHARACTERISTIC CLASSES

Orientation transport

ON COSTELLO S CONSTRUCTION OF THE WITTEN GENUS: L SPACES AND DG-MANIFOLDS

GEOMETRIC QUANTIZATION

LECTURE 28: VECTOR BUNDLES AND FIBER BUNDLES

M4P52 Manifolds, 2016 Problem Sheet 1

The Atiyah bundle and connections on a principal bundle

Chap. 1. Some Differential Geometric Tools

Noether Symmetries and Conserved Momenta of Dirac Equation in Presymplectic Dynamics

GEOMETRIC BRST QUANTIZATION

BRIAN OSSERMAN. , let t be a coordinate for the line, and take θ = d. A differential form ω may be written as g(t)dt,

PREQUANTIZATION OF SYMPLECTIC SUPERMANIFOLDS

1. Geometry of the unit tangent bundle

CONSISTENT INTERACTIONS BETWEEN BF AND MASSIVE DIRAC FIELDS. A COHOMOLOGICAL APPROACH

Reminder on basic differential geometry

The sh Lie structure of Poisson brackets in field theory

New Topological Field Theories from Dimensional Reduction of Nonlinear Gauge Theories

The BRST complex for a group action

The Cohomology of the Variational Bicomplex Invariant under the Symmetry Algebra of the Potential Kadomtsev-Petviashvili Equation

Bulletin of the Transilvania University of Braşov Vol 8(57), No Series III: Mathematics, Informatics, Physics, 43-56

Geometry 9: Serre-Swan theorem

Math 868 Final Exam. Part 1. Complete 5 of the following 7 sentences to make a precise definition (5 points each). Y (φ t ) Y lim

Lie algebra cohomology

Euler-Poincaré reduction in principal bundles by a subgroup of the structure group

Transverse geometry. consisting of finite sums of monomials of the form

Dimensional Reduction of Invariant Fields and Differential Operators. II. Reduction of Invariant Differential Operators

De Rham Cohomology. Smooth singular cochains. (Hatcher, 2.1)

arxiv:alg-geom/ v1 29 Jul 1993

BRST 2006 (jmf) 7. g X (M) X ξ X. X η = [ξ X,η]. (X θ)(η) := X θ(η) θ(x η) = ξ X θ(η) θ([ξ X,η]).

Master of Science in Advanced Mathematics and Mathematical Engineering

ON LIFTS OF SOME PROJECTABLE VECTOR FIELDS ASSOCIATED TO A PRODUCT PRESERVING GAUGE BUNDLE FUNCTOR ON VECTOR BUNDLES

Notes on the geometry of Lagrangian torus fibrations

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

Cohomology jump loci of local systems

arxiv:hep-th/ v2 11 Sep 1996

arxiv:math/ v4 [math.dg] 3 Sep 2006

Hodge theory for bundles over C algebras

Math 231b Lecture 16. G. Quick

Vector fields in the presence of a contact structure

Derivations and differentials

Generalized complex geometry and topological sigma-models

arxiv: v2 [math.dg] 19 Aug 2017

Reduction of Symplectic Lie Algebroids by a Lie Subalgebroid and a Symmetry Lie Group

Wild ramification and the characteristic cycle of an l-adic sheaf

EXTERIOR AND SYMMETRIC POWERS OF MODULES FOR CYCLIC 2-GROUPS

On the Van Est homomorphism for Lie groupoids

A Crash Course of Floer Homology for Lagrangian Intersections

Complex line bundles. Chapter Connections of line bundle. Consider a complex line bundle L M. For any integer k N, let

From symplectic to spin geometry

Let V, W be two finite-dimensional vector spaces over R. We are going to define a new vector space V W with two properties:

QUALIFYING EXAMINATION Harvard University Department of Mathematics Tuesday September 21, 2004 (Day 1)

A geometric solution of the Kervaire Invariant One problem

1 Introduction and preliminaries notions

COMPUTING THE POISSON COHOMOLOGY OF A B-POISSON MANIFOLD

INTEGRATION OF TWISTED POISSON STRUCTURES ALBERTO S. CATTANEO AND PING XU

Math 225B: Differential Geometry, Final

Introduction to finite element exterior calculus

Contact Transformations in Classical Mechanics

Homogeneous para-kähler Einstein manifolds. Dmitri V. Alekseevsky

Higgs Bundles and Character Varieties

Holomorphic line bundles

Vector bundles in Algebraic Geometry Enrique Arrondo. 1. The notion of vector bundle

Homotopy Lie Algebroids

Acceleration bundles on Banach and Fréchet manifolds

Sheaves of Lie Algebras of Vector Fields

COHOMOLOGY OF FOLIATED FINSLER MANIFOLDS. Adelina MANEA 1

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

Spin(10,1)-metrics with a parallel null spinor and maximal holonomy

HARMONIC COHOMOLOGY OF SYMPLECTIC FIBER BUNDLES

COSYMPLECTIC REDUCTION OF CONSTRAINED SYSTEMS WITH SYMMETRY

Lecture I: Constrained Hamiltonian systems

The inverse problem for Lagrangian systems with certain non-conservative forces

Dirac Structures on Banach Lie Algebroids

Hungry, Hungry Homology

A MARSDEN WEINSTEIN REDUCTION THEOREM FOR PRESYMPLECTIC MANIFOLDS

A Problem of Hsiang-Palais-Terng on Isoparametric Submanifolds

THE QUANTUM CONNECTION

Isomorphism classes for Banach vector bundle structures of second tangents. Dodson, CTJ and Galanis, GN and Vassiliou, E. MIMS EPrint: 2005.

Hamilton-Jacobi theory on Lie algebroids: Applications to nonholonomic mechanics. Manuel de León Institute of Mathematical Sciences CSIC, Spain

Dynamical systems on Leibniz algebroids

Some aspects of the Geodesic flow

LECTURE 2. (TEXED): IN CLASS: PROBABLY LECTURE 3. MANIFOLDS 1. FALL TANGENT VECTORS.

Determinant Bundle in a Family of Curves, after A. Beilinson and V. Schechtman

Gauge Fixing and Constrained Dynamics in Numerical Relativity

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

Lie Algebra Cohomology

arxiv:math/ v1 [math.ag] 18 Oct 2003

Transcription:

Noether s second theorem in a general setting. Reducible gauge theories D.Bashkirov 1, G.Giachetta 2, L.Mangiarotti 2, G.Sardanashvily 1 arxiv:math/0411070v1 [math.dg] 3 Nov 2004 1 Department of Theoretical Physics, Moscow State University, Russia 2 Department of Mathematics and Informatics, University of Camerino, Italy Abstract: We prove Noether s direct and inverse second theorems for Lagrangian systems on fiber bundles in the case of gauge symmetries depending on derivatives of dynamic variables of an arbitrary order. The appropriate notions of reducible gauge symmetries and Noether s identities are formulated, and their equivalence is proved. 1 Introduction Different variants of Noether s second theorem state that, if a Lagrangian admits symmetries depending on parameters, its variational derivatives obey certain relations, called Noether s identities. In a rather general setting, this theorem has been formulated in [6]. We present Noether s second theorem and its inverse (Theorem 5.2) for Lagrangian systems on a fiber bundle X in the case of gauge symmetries depending on derivatives of dynamic variables of an arbitrary order. For this purpose, we consider Lagrangian formalism on the composite fiber bundle E X, where E is a vector bundle of gauge parameters. Accordingly, gauge transformations are represented by a linear differential operator υ on E taking its values into the vertical tangent bundle V of X. For instance, if transformations of parameters are independent of fields, E is the pull-back of some vector bundle E X onto. Noether s identity for a Lagrangian L is defined as a differential operator on the fiber bundle (3.9) which takes its values into the density-dual E = E n T X (1.1) of E and whose kernel contains the image of the Euler Lagrange operator δl of L, i.e., δl = 0 (Definition 5.1). Expressed in these terms, Noether s second theorem and its inverse follow at once from the first variational formula (Proposition 4.1) and the properties of differential operators on dual fiber bundles (Theorem 7.1). Namely, there exists the intertwining operator η(υ) = (7.3), η( ) = υ (7.4) such that η(η(υ)) = υ, η(η( )) =, (1.2) η(υ υ ) = η(υ ) η(υ), η( ) = η( ) η( ). (1.3) 1

The appropriate notions of reducible Noether s identities and gauge symmetries are formulated, and their equivalence with respect to the intertwining operator η is proved (Theorem 6.3). A particular attention is paid to global aspects of Noether s second theorem as a preliminary step of the global analysis of BV quantization [7, 9]. Here, we describe classical systems of even fields and gauge transformations. This description however fails to be algebraically self-contained. Firstly, Noether s identity does not determine transformations of gauge parameters because a classical Lagrangian is independent of gauge parameters. Secondly, the weak equality A 0 (i.e., A equals zero on-shell) is not algebraically formulated. In the framework of BV quantization, a Lagrangian depends on ghosts, its gauge symmetry is required to be nilpotent, and Euler Lagrange equations are replaced with the Koszul Tate differential δ acting on antifields [3]. In these terms, the conditions a 0, υ a 0 in Definitions 6.1 and 6.2 of reducible Noether s identities and gauge symmetries mean that the operators a and υ a are not δ-exact. 2 Preliminaries Recall that an r-order Lagrangian on a fiber bundle X is defined as a density L = Lω : J r n T X, ω = dx 1 dx n, n = dim X, (2.1) on the r-order jet manifold J r of sections of X. A k-order differential operator on a fiber bundle X with values into a fiber bundle Z X is defined as a section of the fiber bundle J k Z J k. It admits an m-order X jet prolongation (m) as a section of the fiber bundle J m+k J m Z J m+k. X By a differential operator throughout is meant its appropriate finite order jet prolongation. Given bundle coordinates (x λ, y i ) on and (x λ, z A ) on Z, a differential operator reads z A = A (x λ, y i Λ), z A Σ (m) = d Σ A, 0 Λ k, 0 Σ m, where d Σ are total derivatives (3.4). If Z is a composite fiber bundle π π Z : Z X and the relation π Z = π0 k holds, a differential operator is identified to a section of the fiber bundle J k Z J k or, equivalently, a bundle morphism J k Z. If Z is a vector bundle and 0 is its canonical zero section, Ker is defined as the inverse image of 0. Let W and Z be vector bundles over X. A k-order Z-valued differential operator on W X is called linear on W (or, simply, linear) if : J k W Z is a 2

morphism of the vector bundle J k W J k to the vector bundle Z over π k 0 : Jk. Given bundle coordinates (x λ, y i, w r ) on W and (x λ, y i, z A ) on Z, such an operator reads i = y i, A = 0 Ξ k 3 Lagrangian formalism on fiber bundles A,Ξ r (x λ, yλ i )wr Ξ, 0 Λ k. (2.2) In Lagrangian formalism on a fiber bundle X, we deal with the following graded differential algebra (henceforth GDA). Jet manifolds of X make up the inverse system X π π1 0 J 1 J r 1 πr r 1 J r. (3.1) In the sequel, the index r = 0 stands for. Accordingly, we have the direct system O X π O π1 0 O1 O r 1 πr r 1 Or (3.2) of GDAs O r of exterior forms on jet manifolds Jr with respect to the pull-back monomorphisms πr 1 r. Its direct limit O is a GDA consisting of all exterior forms on finite order jet manifolds modulo the pull-back identification. The projective limit (J, πr : J J r ) of the inverse system (3.1) is a Fréchet manifold [10]. A bundle atlas {(U ; x λ, y i )} of X yields the coordinate atlas {((π 0 ) 1 (U ); x λ, y i Λ)}, y i λ+λ = xµ x λd µy i Λ, 0 Λ, (3.3) of J, where Λ = (λ k...λ 1 ) is a symmetric multi-index, λ + Λ = (λλ k...λ 1 ), and d λ = λ + 0 Λ y i λ+λ Λ i, d Λ = d λr d λ1 (3.4) are the total derivatives. There is the GDA epimorphism O O U, called the restriction of O to the chart (3.3). Then O can be written in a coordinate form where the horizontal one-forms {dx λ } and the contact one-forms {θλ i = dyλ i yλ+λdx i λ } are generating elements of the O 0 U -algebra O U. Though J is not a smooth manifold, the coordinate transformations of elements of O are smooth because these elements are exterior forms on finite order jet manifolds. There is the canonical decomposition O = Ok,m of O into O0 -modules O k,m of k-contact and m-horizontal forms together with the corresponding projectors h k : O Ok, and hm : O O,m. Accordingly, the exterior differential on O is split into the sum d = d H + d V of the nilpotent total and vertical differentials d H (φ) = dx λ d λ (φ), d V (φ) = θ i Λ Λ i φ, φ O. 3

One also introduces the R-module projector = 0<k 1 k h k h n, (φ) = ( 1) Λ θ i [d Λ ( i Λ φ)], 0 Λ φ O >0,n, (3.5) of O such that d H = 0, and the nilpotent variational operator δ = d on O,n. Let us put E k = (O k,n ). Then the GDA O is split into the well-known variational bicomplex [2, 8, 9, 10]. Here, we are concerned with its variational complex 0 R O 0 d H O 0,1 dh O 0,n and the complex of one-contact forms 0 O 1,0 d H O 1,1 dh O 1,n They possess the following cohomology [1, 7, 8]. Theorem 3.1. δ δ E 1 E 2 (3.6) E 1 0. (3.7) (i) The cohomology of the variational complex (3.6) equals the de Rham cohomology of. (ii) The complex (3.7) is exact. Any finite order Lagrangian L (2.1) is an element of O 0,n, while δl = E i θ i ω = ( 1) Λ d Λ ( i Λ L)θi ω E 1 (3.8) 0 Λ is its Euler Lagrange operator taking the values into the vector bundle T ( n T X) = V Then item (i) of Theorem 3.1 states the following. Theorem 3.2. Every δ-closed Lagrangian L O 0,n is the sum n T X. (3.9) L = h 0 ψ + d H σ, σ O 0,n 1, (3.10) where ψ is a closed n-form on. Item (ii) of Theorem 3.1 provides the R-module decomposition O 1,n = E 1 d H (O 1,n 1 ). Given a Lagrangian L O 0,n, we have the corresponding decomposition dl = δl d H Ξ (3.11) where Ξ L = Ξ + L is a Lepagean equivalent of L. This decomposition leads to the first variational formula (4.4). 4

Given a Lagrangian L and its Euler-Lagrange operator δl (3.8), we further abbreviate A 0 with an equality which holds on-shell. This means that A is an element of a module over the ideal I L of the ring O 0 which is locally generated by the variational derivatives E i and their total derivations d Λ E i. In this case A vanishes on the kernel of an appropriate jet prolongation of the Euler Lagrange operator. However, the converse need not be true, unless this kernel is an imbedded submanifold. We say that I L is a differential ideal because, if a local functions f belongs to I L, then every total derivative d Λ f does as well. 4 Gauge symmetries in a general setting Let do 0 be the O 0 -module of derivations of the R-ring O 0. Any ϑ do 0 yields the graded derivation (the interior product) ϑ φ of the GDA O given by the relations ϑ df = ϑ(f), f O 0, ϑ (φ σ) = (ϑ φ) σ + ( 1) φ φ (ϑ σ), φ, σ O, and its derivation (the Lie derivative) L ϑ φ = ϑ dφ + d(ϑ φ), φ O, (4.1) L ϑ (φ φ ) = L ϑ (φ) φ + φ L ϑ (φ ). A derivation ϑ do 0 is called vertical if it vanishes on the subring C (X) O 0 of smooth functions on X. Hereafter, only vertical derivations are considered. A derivation ϑ is called contact if the Lie derivative L ϑ (4.1) preserves the contact ideal of the GDA O. Relative to an atlas (3.3), a contact derivation is given by the expression ϑ = 0 Λ d Λ υ i Λ i, (4.2) where { i Λ} is the dual to the basis {θi Λ } with respect to the interior product [9]. A contact derivation is completely determined by its summand υ = υ i (x λ, y i Λ ) i, 0 Λ k, (4.3) which is also a derivation of O 0. It is a section of the pull-back V J k J k of the vertical tangent bundle V onto J k [4] and, thus, is a k-order V -valued differential operator on. Accordingly, ϑ (4.2) is its infinite order jet prolongation. One calls υ (4.3) a generalized vector field on. Proposition 4.1. It follows from the splitting (3.11), that the Lie derivative of a Lagrangian L (2.1) along a contact derivation ϑ (4.2) admits the canonical decomposition L ϑ L = υ δl + d H (h 0 (ϑ Ξ L )), (4.4) 5

where Ξ L is a Lepagean equivalent of L [9]. A generalized vector field υ (4.3) is called a variational symmetry of a Lagrangian L if the Lie derivative L ϑ L (or, equivalently, υ δl) is a d H -exact form. A Lagrangian system on a fiber bundle X is said to be a gauge theory if its Lagrangian L admits a family of variational symmetries parameterized by elements of a vector bundle E as follows. Remark 4.1. In a general setting, one can consider a vector bundle E over some finite jet order manifold J m of. However, any vector bundle E J m is the pull-back onto J m of a vector bundle over because is a strong deformation retract of J m. Let E be a vector bundle coordinated by (x λ, y i, ξ r ). Given a Lagrangian L on, let us consider its pull-back, say again L, onto E. Let ϑ E be a contact derivation of the R-ring O 0 E, whose restriction ϑ = ϑ E O 0 = 0 Λ d Λ υ i Λ i (4.5) to O 0 O0 E is linear in coordinates ξr Ξ. It is determined by a generalized vector field υ E on E whose canonical projection υ : J k E υ E V E E V (see the exact sequence (4.8) below) is a linear V -valued differential operator υ = 0 Ξ m υ i,ξ r (x λ, y i Σ )ξr Ξ i (4.6) on E. Let ϑ E be a variational symmetry of a Lagrangian L on E, i.e., υ E δl = υ δl = d H σ. (4.7) Then one says that υ (4.6) is a gauge symmetry of a Lagrangian L. This is the case of Noether s second theorem. Note that any generalized vector field υ (4.6) gives rise to a generalized vector field υ E on E and, thus, defines a contact derivation ϑ E of O 0 E. Indeed, let us consider the exact sequence of vector bundles 0 V E V E E V 0, (4.8) where V E is the vertical tangent bundle of E. Its splitting Γ lifts υ to the generalized vector field υ E = Γ υ on E. However, the Lie derivative L ϑe L = υ δl + d H (h 0 (ϑ Ξ L )) 6

depends only on υ, but not a lift Γ. Remark 4.2. One can also consider a local-variational gauge symmetry ϑ E when the Lie derivative L ϑe L is δ-closed, but any local-variational gauge symmetry is variational as follows. By virtue of Theorem 3.2, L ϑe L takes the form (3.10) where ψ is a closed form on E. Since E is a vector bundle, is a strong deformation retract of E. Then any non-exact closed form on E is the pull-back of that on, i.e., it is independent on fiber coordinates ξ r. Since the Lie derivative L ϑ L is linear in ξλ, r it is always d H -exact, i.e., ϑ E is variational. 5 Noether s second theorem Let us start with the notion of Noether s identity. Definition 5.1. Given a Lagrangian L (2.1), we say that its Euler Lagrange operator δl (3.8) obeys Noether s identity if there exists a vector bundle E and a linear differential operator of order 0 m on the vector bundle (3.9) with the values into the density-dual E (1.1) of E such that δl = 0. (5.1) Given bundle coordinates (x λ, y i, y i ) on the fiber bundle (3.9) and (x λ, y i, ξ r ) on E, a differential operator in Definition 5.1 is represented by the density = r ξ r ω = i,λ r (x λ, yσ)y j Λi ξ r ω O 0,n [E V ], 0 Σ m. (5.2) Then Noether s identity (5.1) takes the coordinate form Remark 5.1. We will use the relations 0 Λ k 0 Λ k η(b) Λ = B Λ d Λ A = 0 Λ k ( 1) Λ d Λ (B Λ A) = 0 Σ k Λ i,λ r d Λ E i ξ r ω = 0. (5.3) ( 1) Λ d Λ (B Λ )A + d H σ, (5.4) 0 Λ k η(b) Λ d Λ A, (5.5) ( 1) Σ+Λ C Σ Σ+Λ d ΣB Σ+Λ, Cb a = b! a!(b a)!, (5.6) 7

for any exterior form A O Z and any local function BΛ O 0 Z on jet manifolds of a fiber bundle Z X. Since k ( 1) a Ck a = 0 for k > 0, it is easily verified that a=0 (η η)(b) Λ = B Λ. (5.7) Theorem 5.2. If a Lagrangian L (2.1) admits a gauge symmetry υ (4.6), its Euler Lagrange operator obeys Noether s identity (5.3) where i,λ r = η(υ) i,λ r = 0 Σ m Λ ( 1) Σ+Λ C Σ Σ+Λ d Συr i,σ+λ. (5.8) Conversely, if the Euler Lagrange operator of a Lagrangian L obeys Noether s identity (5.3), this Lagrangian admits a gauge symmetry υ (4.6) where υ i,λ r = η( ) i,λ r = 0 Σ m Λ ( 1) Σ+Λ C Σ Σ+Λ d Σ i,σ+λ r. (5.9) Proof. Given an operator υ (4.6), the operator = η(υ) expressed in the coordinate form (5.8) is defined in accordance with Theorem 7.1. Since the density is d H -exact, Noether s identity υ δl = υ i E i ω = 0 Ξ m υ i,ξ r ξ r Ξ E iω δ(υ δl) = η(υ) δl = 0 holds. Conversely, any operator (5.2) defines the generalized vector field υ = η( ) expressed in the coordinate form (5.9). Due to Noether s identity (5.3), we obtain 0 = 0 Ξ m ξ r i,λ r d Λ E i ω = i.e., υ is a gauge symmetry of L. υ i,ξ r ξ r Ξ E iω + d H σ = υ δl + dσ H, ( 1) Λ d Λ (ξ r i,λ r )E i ω + d H σ = By virtue of the relations (1.2), there is one-to-one correspondence between gauge symmetries of a Lagrangian L and Noether s identities for δl. Any Lagrangian L have gauge symmetries. In particular, there always exist trivial gauge symmetries υ = Λ Tr j,i,λ E j ξλ r i, Tr j,i,λ = Tr i,j,λ. 8

Furthermore, given a gauge symmetry υ (4.6), let h be a linear differential operator on some vector bundle E, coordinated by (x λ, y i, ξ s ), with values into the vector bundle E. Then the composition υ 0 = υ h = υ i,λ s ξ s Λ i, υ i,λ s = Ξ+Ξ =Λ 0 Σ m Ξ υr i,ξ+σ d Σ h r,ξ s, is a gauge symmetry of the pull-back of a Lagrangian L onto E. In view of this ambiguity, we agree to say that a gauge symmetry υ (4.6) of a Lagrangian L is complete if any different gauge symmetry υ 0 of L factorizes through υ as υ 0 = υ h + T, T 0. A complete gauge symmetry always exists, but the vector bundle of its parameters need not be finite-dimensional. Accordingly, given Noether s identity (5.1), let H be a linear differential operator on E with values into the density-dual E (1.1) of some vector bundle E. Then the composition = H also provides Noether s identity for δl. We agree to call Noether s identity (5.1) complete if any different differential operator possessing this property factorizes through as = H + F, F 0. Proposition 5.3. A gauge symmetry υ of a Lagrangian L is complete iff so is associated Noether s identity. Proof. The proof follows at once from Proposition 7.2. Given a gauge symmetry υ of L, let υ 0 be a different gauge symmetry. If η(υ) provides complete Noether s identity, then η(υ 0) = H η(υ) + F, F 0, and, by virtue of the relations (1.3), we have υ 0 = υ η(h) + η(f), where η(f) 0 because I L is a differential ideal. The converse is similarly proved. 6 Reducible gauge theories Recall that the notion of reducible Noether s identity has come from that of a reducible constraint [5], but it involves differential relations. 9

Definition 6.1. Complete Noether s identity (5.1) with 0 is said to be k-stage reducible (k = 0, 1,...) if there exist vector bundles E a and differential operators a, a = 0,..., k, such that: (i) a is a linear differential operator on the density-dual E a 1 of E a 1 with values into the density-dual E a of E a, where E 1 = E; (ii) a 0 for all a = 0,..., k; (iii) a a 1 0 for all a = 0,..., k, where 1 = ; (iv) if a is another differential operator possessing these properties, then it factorizes through a on-shell. In particular, a zero-stage reducible Noether s identity is called reducible. In this case, given bundle coordinates (x λ, y i, ξ r ) on E and (x λ, y i, ξ r 0 ) on E 0, a differential operator 0 reads 0 = 0 Ξ m 0 r,ξ r 0 ξ Ξr ξ r 0 ω. (6.1) Then the reduction condition 0 0 for Noether s identity (5.3) reads 0 Ξ m 0 r,ξ r 0 d Ξ ( i.e., the left hand-side of this expression takes the form 0 Σ m 0 +m i,λ r y Λi )ξ r 0 ω 0, (6.2) M i,σ r 0 y Σi ξ r 0 ω, where all the coefficients Mr i,σ 0 belong to the ideal I L. Given Noether s identity (5.3), one can think of the reduction condition (6.2) as being an equation for the local functions r,ξ r 0. If all its solutions vanish on-shell, Noether s identity is irreducible. Definition 6.2. A complete gauge symmetry υ 0 (4.6) is said to be k-stage reducible if there exist vector bundles E a and differential operators υ a, a = 0,..., k, such that: (i) υ a is a linear differential operator on the vector bundle E a with values into the vector bundle E a 1 ; (ii) υ a 0 for all a = 0,...,k; (iii) υ a 1 υ a 0 for all a = 0,...,k, where υ 1 = υ; (iv) if υ a is another differential operator possessing these properties, then υ a factorizes through υ a on-shell. Theorem 6.3. A gauge symmetry υ is k-stage reducible iff so is the associated Noether s identity. 10

Proof. The proof follows at once from Theorem 7.1 and Proposition 7.2. Let us put a = η(υ a ), a = 0,...,k. If υ a 0, then η(υ a ) 0 because I L is a differential ideal. By the same reason, if υ a 1 and υ a obey the reduction condition υ a 1 υ a 0, then η(υ a 1 υ a ) = η(υ a ) η(υ a 1 ) 0. The converse is justified in the same way. The equivalence of the conditions in items (iv) of Definitions 6.1 and 6.2 is proved similarly to that in Proposition 5.3 7 Appendix. Differential operators on dual fiber bundles Given a fiber bundle X, let E and Q be vector bundles coordinated by (x λ, y i, ξ r ) and (x λ, y i, q a ), respectively. Let E and Q be their density-duals (1.1) coordinated by (x λ, y i, ξ r ) and (x λ, y i, q a ), respectively. Let υ be a linear Q-valued differential operator of order m on E. It is represented by the function υ = υ a q a = υr a,λ (x λ, yσ i )ξr Λ q a O 0 [E Q ], 0 Σ m. (7.1) Let be an m-order linear differential operator on the density-dual Q of Q with values into the density-dual E of E. It is represented by the density = r ξ r ω = a,λ r (x λ, y i Σ )q Λaξ r ω O 0,n [E Q ], 0 Σ m, (7.2) Theorem 7.1. E -valued differential operator Any linear Q-valued differential operator υ (7.1) on E yields the linear η(υ) = η(υ) a,λ r = η(υ) a,λ r q Λa ξ r ω, (7.3) 0 Σ m Λ ( 1) Σ+Λ C Σ Σ+Λ d Σ(υr a,σ+λ ), on Q. Conversely, any linear E -valued differential operator (7.2) on Q defines the linear Q-valued differential operator η( ) = η( ) a,λ r = on E. The relations (1.2) hold. η( ) a,λ r ξ r Λ q a, (7.4) 0 Σ m Λ ( 1) Σ+Λ C Σ Σ+Λ d Σ( a,σ+λ r ), 11

Proof. The function υ (7.1) defines the density Its Euler Lagrange operator υ = takes its values into the fiber bundle where V (E υ a,λ r ξ r Λ q aω O 0,n [E δ(υ) = E i dy i ω + E r dξ r ω + E a dq a ω V (E Q ) E Q n Q ]. (7.5) T X, (7.6) Q ) is the vertical cotangent bundle of the fiber bundle E Q X. Let α E : V (E Q ) V (E Q ) V E (7.7) be the canonical projection of V (E Q ) onto the vertical cotangent bundle V (E Q ) of the fiber bundle E Q and, afterwards, onto the vertical cotangent bundle V E of E. Then we obtain a differential operator (α E δ)(υ) on E Q with values into the fiber bundle V E n T X. It reads E (α E δ)(υ) = E r dξ r ω = ( 1) Λ d Λ (υr a,λ q a )dξ r ω, where {dξ r } is the fiber basis for V E E. Due to the canonical isomorphism V E = E E, this operator defines the density ( 1) Λ d Λ (υr a,λ q a )ξ r ω O 0,n [E Q ] and, by virtue of the formula (5.5), the desired differential operator (7.3). Conversely, the Euler Lagrange operator of the density (7.2) takes values into the fiber bundle (7.6) and reads δ(υ) = E i dy i ω + E r dξ r ω + E a dq a ω. (7.8) In order to repeat the above mentioned procedure, let us consider a volume form Jω on X and substitute dq a ω = Jdq a ω into the expression (7.8). Using the projection α Q : V (E Q ) V Q similar to α E (7.7) and the canonical isomorphism V Q = Q Q, we come to the density ( 1) Λ d Λ ( a,λ r ξ r )q a Jω O 0,n [E Q ] 12

and, hence, the function ( 1) Λ d Λ ( a,λ r ξ r )q a O 0 [E Q ], defining the desired operator (7.4). The relations (1.2) result from the relation (5.7). Relations (1.2) show that the intertwining operator η (7.3) (7.4) provides a bijection between the sets Diff(E, Q) and Diff(Q, E ) of differential operators (7.1) and (7.2). Proposition 7.2. Compositions of operators υ υ and obey the relations (1.3). Proof. It suffices to prove the first relation. Let υ υ Diff(E, Q) be a composition of differential operators υ Diff(E, Q) and υ Diff(E, E). Given fibered coordinate (ξ r ) on E, (ǫ p ) on E and (q a ) on Q, this composition defines the density (7.5) υ υ = Λ υr a,λ d Λ ( Σ υ r,σ p ǫ p Σ)q a ω. Following the relation (5.4), one can bring this density into the form Σ υ p r,σ ǫ p Σ Λ ( 1) Λ d Λ (υ a,λ r q a )ω + d H σ = Σ υ p r,σ ǫ p Σ Λ η(υ) a,λ r q Λa ω + d H σ. Its Euler Lagrange operator projected to V E E T X is Σ ( 1) Σ d Σ (υ r,σ p Λ η(υ) a,λ r q Λa )dǫ p ω = Σ η(υ ) r,σ p d Σ( Λ η(υ) a,λ r q Λa )dǫ p ω, that leads to the desired composition η(υ ) η(υ). References [1] Anderson, I. and Duchamp, T.: On the existence of global variational principles. Amer. J. Math. 102, 781-868 (1980) [2] Anderson, I.: Introduction to the variational bicomplex. Contemp. Math. 132, 51-73 (1992) [3] Barnich, G., Brandt, F. and Henneaux, M.: Local BRST cohomology in gauge theories. Phys. Rep. 338, 439-569 (2000) [4] Fatibene, L., Ferraris, M., Francaviglia, M. and McLenaghan, R.: Generalized symmetries in mechanics and field theories. J. Math.. Phys. 43, 3147-3161 (2002) 13

[5] Fish, M. and Henneaux, M.: Homological perturbation theory and algebraic structure of the antifield-antibracket formalism for gauge theories. Commun. Math. Phys. 128 627-640 (1990) [6] Fulp, R., Lada, T. and Stasheff, J.: Noether variational Theorem II and the BV formalism. Rend. Circ. Mat. Palermo (2) Suppl. No. 71, 115-126 (2003); E-print arxiv: math.qa/0204079 [7] Giachetta, G., Mangiarotti, L. and Sardanashvily, G.: Iterated BRST cohomology. Lett. Math. Phys. 53, 143-156 (2000) [8] Giachetta, G., Mangiarotti, L. and Sardanashvily, G.: Cohomology of the infinite-order jet space and the inverse problem. J. Math. Phys. 42, 4272-4282 (2001) [9] Giachetta, G., Mangiarotti, L. and Sardanashvily, G.: Lagrangian supersymmetries depending on derivatives. Global analysis and cohomology. Commun. Math. Phys. (accepted); E-print arxiv: hep-th/0407185 [10] Takens, F.: A global version of the inverse problem of the calculus of variations. J. Diff. Geom. 14, 543-562 (1979) 14